NHC-Catalyzed Reactions of Aryloxyacetaldehydes: A Domino

(12) The likely pathway involves initial addition of the NHC to aldehyde 1 (Scheme 2).(13) A formal 1 .... (c) Arya , P.; Qin , H. P. Tetrahedron 2000...
0 downloads 0 Views 194KB Size
ORGANIC LETTERS

NHC-Catalyzed Reactions of Aryloxyacetaldehydes: A Domino Elimination/Conjugate Addition/Acylation Process for the Synthesis of Substituted Coumarins

2009 Vol. 11, No. 1 105-108

Eric M. Phillips, Manabu Wadamoto, Howard S. Roth, Andrew W. Ott, and Karl A. Scheidt* Department of Chemistry, Northwestern UniVersity, EVanston, Illinois 60208 [email protected] Received October 23, 2008

ABSTRACT

N-Heterocyclic carbenes (NHCs) catalyze a domino Michael addition/acylation reaction to form 3,4-dihydrocoumarins. The reaction proceeds through addition of the NHC to an aryloxyaldehyde followed by elimination of a phenoxide leaving group, generating an enol intermediate. This transient nucleophile generated in situ performs a 1,4-addition onto a conjugate acceptor, and the carbene catalyst is regenerated upon acylation of the phenoxide anion resulting in formation of 3,4-dihydrocoumarins.

Enolates are essential intermediates in chemistry and biology due to their high utility in carbon-carbon bond-forming reactions.1 The catalytic, in situ generation of enolates or enols has become increasingly important due to their efficiency in synthesis and emerging environmental concerns. Consequently, significant effort in this area has been directed toward metal-catalyzed processes2 that allow for the addition (1) (a) Evans, D. A. Aldrichimica Acta 1982, 15, 23. (b) Mukaiyama, T.; Kobayashi, S. Org. React. (N.Y.) 1994, 46, 1–103. (c) Arya, P.; Qin, H. P. Tetrahedron 2000, 56, 917–947. (2) (a) Yoshikawa, N.; Yamada, Y. M. A.; Das, J.; Sasai, H.; Shibasaki, M. J. Am. Chem. Soc. 1999, 121, 4168–4178. (b) Trost, B. M.; Ito, H. J. Am. Chem. Soc. 2000, 122, 12003–12004. (c) Hamashima, Y.; Hotta, D.; Sodeoka, M. J. Am. Chem. Soc. 2002, 124, 11240–11241. (d) Evans, D. A.; Thomson, R. J. J. Am. Chem. Soc. 2005, 127, 10506–10507. (e) Mohr, J. T.; Behenna, D. C.; Harned, A. M.; Stoltz, B. M. Angew. Chem., Int. Ed. 2005, 44, 6924–6927. (f) Trost, B. M.; Bream, R. N.; Xu, J. Angew. Chem., Int. Ed. 2006, 45, 3109–3112. 10.1021/ol802448c CCC: $40.75 Published on Web 12/02/2008

 2009 American Chemical Society

of the corresponding metalloenolate to various electrophiles.3 More recently, attention has turned toward the formation of enolates using only organic molecules as catalysts.4 Numerous groups5-8 have employed secondary amines to catalyze enantioselective aldol, Michael, and Mannich reactions (3) Sodeoka, M.; Hamashima, Y. Bull. Chem. Soc. 2005, 78, 941–956. (4) Mukherjee, S.; Yang, J. W.; Hoffmann, S.; List, B. Chem. ReV. 2007, 107, 5471–5569. (5) (a) List, B.; Lerner, R. A.; Barbas, C. F. J. Am. Chem. Soc. 2000, 122, 2395–2396. (b) Fonseca, M. T. H.; List, B. Angew. Chem., Int. Ed. 2004, 43, 3958–3960. (6) Brown, S. P.; Brochu, M. P.; Sinz, C. J.; MacMillan, D. W. C. J. Am. Chem. Soc. 2003, 125, 10808–10809. (7) Hayashi, Y.; Yamaguchi, J.; Sumiya, T.; Shoji, M. Angew. Chem., Int. Ed. 2004, 43, 1112–1115. (8) (a) Momiyama, N.; Torii, H.; Saito, S.; Yamamoto, H. Proc. Natl. Acad. Sci. U.S.A. 2004, 101, 5374–5378. (b) Yamamoto, Y.; Momiyama, N.; Yamamoto, H. J. Am. Chem. Soc. 2004, 126, 5962–5963.

through the generation of a reactive enamine intermediate.9 We report here that N-heterocyclic carbenes (NHCs) catalyze the formation of enolates/enols through an elimination process of R-aryloxy aldehydes. The capture of these nucleophiles leads to substituted coumarins by a domino Michael addition/acylation process (eq 1 in Scheme 1).

Scheme 2. Proposed Reaction Pathway

Scheme 1. General Concept

In our recent carbene catalysis studies to access enols by the protonation of homoenolates,10 we became interested in whether new carbon-carbon bond-forming reactions were possible using reactive acetate-type enol-NHC intermediates from R-aryloxyacetaldehydes. This concept is distinctive since the phenol in the R-position serves an important dual purpose: first to facilitate generation of the enol equivalent and second to liberate the carbene catalyst through an acylation event.11 We hypothesized that this process could be evaluated with an aldehyde containing a tethered conjugate acceptor in the ortho-position.12 The likely pathway involves initial addition of the NHC to aldehyde 1 (Scheme 2).13 A formal 1,2-shift of the aldehyde proton results in elimination of phenoxide/phenol derivative III and concomitant formation of enol equivalent II. A Michael addition could then occur which, when followed by intramolecular O-acylation, promotes regeneration of the carbene catalyst. A delicate balance of leaving group ability and nucleophilicity of the R-position substituent is crucial. A catalytic process is untenable if these factors are not optimal in this new type of “rebound” process. For example, a strong leaving (9) List, B. Acc. Chem. Res. 2004, 37, 548–557. (10) (a) Phillips, E. M.; Wadamoto, M.; Chan, A.; Scheidt, K. A. Angew. Chem., Int. Ed. 2007, 46, 3107–3110. (b) Wadamoto, M.; Phillips, E. M.; Reynolds, T. E.; Scheidt, K. A. J. Am. Chem. Soc. 2007, 129, 10098-10099. For excellent reviews on carbene catalysis, see: (c) Enders, D.; Niemeier, O.; Henseler, A. Chem. ReV. 2007, 107, 5606–5655. (d) Marion, N.; DiezGonzalez, S.; Nolan, I. P. Angew. Chem., Int. Ed. 2007, 46, 2988–3000. (11) For a review on domino reactions, see: (a) Tietze, L. F. Chem. ReV. 1996, 96, 115-136. For NHC reactions using unstable a-chloroaldehydes in which the leaving group (chloride) is not involved in catalysts turnover, see: (b) Reynolds, N. T.; Rovis, T. J. Am. Chem. Soc. 2005, 127, 16406–16407. (c) He, M.; Uc, G. J.; Bode, J. W. J. Am. Chem. Soc. 2006, 128, 15088–15089. (12) (a) Yang, J. W.; Chandler, C.; Stadler, M.; Kampen, D.; List, B. Nature 2008, 452, 453–455. (b) Hayashi, Y.; Itoh, T.; Ohkubo, M.; Ishikawa, H. Angew. Chem., Int. Ed. 2008, 47, 4722–4724. (13) Reynolds, T. E.; Stern, C. A.; Scheidt, K. A. Org. Lett. 2007, 9, 2581–2584. 106

group (e.g., chloride or sulfonate) would potentially initiate the reaction faster (I to II and III) but would fail to turn over the catalyst due to weak nucleophilicity (IV to 2). Enone 1a was chosen to explore this carbene-catalyzed enolate formation strategy and leads to the formation of 3,4dihydrocoumarin products.14 Substituted coumarins and related compounds are biologically active,15 and access to this particular substitution pattern by Michael additions to the parent coumarin remains challenging.16 For example, to the best of our knowledge, there are no general and highyielding reports of conjugate additions of enolates to coumarins. While combining enone 1a with imidazolium or benzimidazolium salts in the presence of base produced a minor amount of the desired product (entries 1 and 2, Table 1), a large increase in yield was observed when triazolium precatalyst C17 was used with DBU (entry 5). Further optimization revealed that acetonitrile was the best solvent (60% yield, entry 6). In this particular case, the moderate yield was due to decomposition from prolonged exposure to silica gel. Indeed, quick filtration through SiO2 provided an improved 83% yield of pure desired product 2a (entry 7). With azolium salt C identified as the most efficient precatalyst to promote this well-orchestrated domino process, we investigated the scope of the reaction. Electron-withdrawing (Table 2, entry 3) and electron-donating aryl ketones (entries 4 and 5), as well as naphthyl ketones (entry 6), are accommodated by the reaction conditions. Electron-with(14) (a) Borges, F.; Roleira, F.; Milhazes, N.; Santana, L.; Uriarte, E. Curr. Med. Chem. 2005, 12, 887–916. (b) Murray, R. D. H. Nat. Prod. Rep. 1995, 12, 477–505. (15) (a) Okuda, T.; Kimura, Y.; Yoshida, T.; Hatano, T.; Okuda, H.; Arichi, S. Chem. Pharm. Bull. 1983, 31, 1625–1631. (b) Demir, A. S.; Gross, R. S.; Dunlap, N. K.; Bashirhashemi, A.; Watt, D. S. Tetrahedron Lett. 1986, 27, 5567–5570. (c) Posakony, J.; Hirao, M.; Stevens, S.; Simon, J. A.; Bedalov, A. J. Med. Chem. 2004, 47, 2635–2644. (16) (a) Connor, R.; McClellan, W. R. J. Org. Chem. 1939, 3, 570– 577. (b) Dixon, D. J.; Horan, R. A. J.; Monck, N. J. T.; Berg, P. Org. Lett. 2004, 6, 4427–4429. (c) Matsuda, T.; Shigeno, M.; Murakami, M. J. Am. Chem. Soc. 2007, 129, 12086–12087. (17) Chan, A.; Scheidt, K. A. J. Am. Chem. Soc. 2008, 130, 2740–2741. Org. Lett., Vol. 11, No. 1, 2009

Table 1. Optimization of Conditions

a

Table 2. Substrate Scope

All reactions performed on 0.2 mmol scale. b Isolated yields.

drawing and -donating groups (entries 7-12) can also be placed on the aromatic ring in a variety of patterns. The 3,4dihydrocoumarin derived from substrate 7 (entry 11) is formed in good yield, although the diastereoselectivity with achiral NHCs is marginal. Additionally, catalyst loading can be decreased to 5 mol % without diminishing the yield (entry 2). Currently, substitution at the R-position of the aldehyde is not accommodated under the reaction conditions. The use of saturated substrates, which replace the phenoxide leaving group with an alkoxide, results in no reaction. From these data, it seems that normal alcohols are not sufficiently stabilized to undergo the first step of the catalytic cycle, i.e., elimination from the initial carbene·aldehyde adduct I. An extensive survey of known chiral azolium salts as catalysts in this reaction provides products in moderate yields but low levels of enantioselectivity (